pH-Sensitive C–ON Bond Homolysis of Alkoxyamines of Imidazoline

Jun 2, 2011 - Copyright © 2011 American Chemical Society. E-mail: [email protected]; [email protected]. Cite this:J. Org. Chem. 76, 14, 5558-5573 ...
0 downloads 0 Views 2MB Size
ARTICLE pubs.acs.org/joc

pH-Sensitive CON Bond Homolysis of Alkoxyamines of Imidazoline Series with Multiple Ionizable Groups As an Approach for Control of Nitroxide Mediated Polymerization Mariya V. Edeleva,*,†,‡ Igor A. Kirilyuk,§ Irina F. Zhurko,§ Dmitry A. Parkhomenko,†,‡ Yuri P. Tsentalovich,† and Elena G. Bagryanskaya*,† †

International Tomography Center SB RAS, Institutskaya 3A, Novosibirsk, 630090, Russian Federation Novosibirsk State University, Pirogova 2, Novosibirsk, 630090, Russian Federation § Novosibirsk Institute of Organic Chemistry SB RAS, Pr. Lavrentjeva 16, Novosibirsk, 630090, Russian Federation ‡

bS Supporting Information ABSTRACT:

Recently, a new concept of pH-switchable agents for reversible additionfragmentation chain transfer (RAFT) polymerization has been introduced by Benaglia et al. (J. Am. Chem. Soc. 2009, 131, 69146915). In this paper we extended the concept of pH-switchable mediators to nitroxide mediated polymerization (NMP) by employing nitroxides with basic or acidic groups as controlling agents. Four alkoxyamines, the derivatives of 2-(4-(dimethylamino)-2-ethyl-5,5-dimethyl-2-(pyridin-4-yl)-2,5-dihydro1H-imidazol-1-oxyl and 2-(2-carboxyethyl)-5,5-diethyl-2,4-dimethyl-2,5-dihydro-1H-imidazol-1-oxyl, have been prepared. The influence of pH on alkoxyamine homolysis rate constants (kd) and on the nitroxide-alkyl radical recombination rate constants (kc) was studied. All alkoxyamines under study as well as the parent nitroxides have several basic groups, which under pH variation can undergo consecutive protonation. It was shown that the kd value under basic conditions are significantly (up to 15-fold) higher than in acidic solution at the same temperature, whereas the kc value in basic solutions decrease by a factor of 2 only. The efficiency of NMP is known to be dependent on kd and kc, both constants being dependent on the monomer structure; therefore the performance of NMP of different monomers in the controlled mode requires different conditions. It is shown that the pH value crucially affects the polymerization regime, changing it from the controlled to the uncontrolled mode. The controlled regime of NMP of different hydrophilic monomers (sodium 4-styrenesulphonate and acrylamide) in aqueous solution under mild conditions (90 °C) can be achieved using the same alkoxyamine by the variation of the pH value. The chain length of polymers depends on pH value during the polymerization.

’ INTRODUCTION Nitroxide mediated radical polymerization (NMP) is a versatile technique for the preparation of homo- and co-polymers with controlled molecular weight, complex architecture, and reactive end groups.17 The analysis of NMP kinetics shows that the polymerization efficiency depends on the rate constant kd of the r 2011 American Chemical Society

dormant species (alkoxyamines) homolysis and on the rate constant of nitroxyl and alkyl radicals recombination kc (Scheme 1).8 The rate constants kd and kc depend on the monomer structure, Received: March 2, 2011 Published: June 02, 2011 5558

dx.doi.org/10.1021/jo200341m | J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

ARTICLE

Scheme 1. Simplified Reaction Scheme of NMP

Chart 1. Structures of Compounds under Investigation

therefore the polymerization of various monomers requires different conditions to perform NMP in the controlled mode.912 Recently, a new concept of pH-switchable agents for reversible additionfragmentation chain transfer (RAFT) polymerization has been introduced.13,14 It was shown that the activation/ deactivation parameters of the RAFT agents can ary upon the protonation of functional groups in the structure of dithiocarbamates used as control agents. Using this effect, the polymerization of so-called “more activated monomers” and “less activated monomers” can be performed in a controlled manner with the same initiator/controlling agent. This opens a facile route to the preparation of block co-polymers. Marx et al.15 used N-tert-butylN-[(1-(2-pyridyl)-2-methyl)-propyl] nitroxide (pyridyl-modified TIPNO) for the polymerization of styrene and acrylates and noted that alkoxyamines with the protonable groups could be of interest for NMP. To describe their unexpected results of polymerization, Mazarin et al.16 proposed that the protonation of the nitroxides moiety can decrease kd value dramatically. However, no experimental evidence of the effect of protonation on kd value was provided. In this paper we extended the concept of pH-switchable mediators to NMP by employing nitroxides with basic or acidic groups as controlling agents. It is known17 that the variation of kd and kc values can significantly affect the polymerization. It was shown that the values of kd and kc strongly depend on the structure of substituents.18,19 The presence of electron-withdrawing groups in the structure of nitroxyl fragment results in the increase of the activation energy of alkoxyamine homolysis and, therefore, in the decrease of the homolysis rate constant kd.18,20 The reversible protonation of functional groups in alkoxyamine changes the polarity of the nitroxyl or alkyl moieties of the alkoxyamine. Thus, it is expected that the value of kd should be pH-dependent. The dependence of kd on the protonation of alkyl moiety of alkoxyamine was recently observed by Bremond et al.21 One can also expect the changes of kc due to the reversible protonation of functional groups in nitroxide. Variation of pH causes strong changes in the EPR spectra of imidazoline nitroxides22 due to the proximity of the basic center at the nitrogen atom N3 (Chart 1) of the heterocycle to the nitroxide group. Introduction of yet another basic center into the side chain can further increase the effect of pH on EPR spectra due to the consecutive protonation of basic groups.23,24 One could expect that the cross-coupling rate constants kc for these nitroxides and kd for corresponding alkoxyamines with several basic groups would also demonstrate the strongest dependence on pH, with the multistep “switching” (variation) upon pH changes. This may permit the regulation of the parameters for the various polymerization processes via pH variation. The efficiency of imidazoline nitroxides as polymerization mediators for acrylates and styrene has recently been demonstrated.2528 The controlled polymerization of methyl

methacrylate (MMA) up to moderate conversion was performed in the presence of imidazoline-based alkoxyamine; however, the polymerization was not living due to the instability of the mediator.29 Since imidazoline nitroxides are water-soluble, they could be used as mediators of NMP of hydrophilic monomers. In most cases, the polymerization of hydrophilic monomers by controlled techniques requires the chemical protection of functional groups,30 and the polymerization temperature has to exceed the water boiling point.31 Only recently, the development of new types of nitroxides made it possible to conduct NMP in aqueous solutions at temperatures as low as 8595 °C.32,33 In this work, we present the first study of the effect of pH on the rate constants kd and kc and the first example of the application of this effect to the polymerization of water-soluble monomers. Four alkoxyamines, the derivatives of 2-(4-(dimethylamino)-2-ethyl5,5-dimethyl-2-(pyridin-4-yl)-2,5-dihydro-1H-imidazol-1-oxyl (1) and 2-(2-carboxyethyl)-5,5-diethyl-2,4-dimethyl-2,5-dihydro-1Himidazol-1-oxyl (2) were synthesized. The pH-dependent reversible protonation of basic functional groups of alkoxyamines 36 and nitroxide 2 was studied using NMR and EPR, correspondingly. The homolysis rate constants for alkoxyamines 36 (Chart 1) based on nitroxides 1 and 2 and the rate constants kc for the nitroxides 1 and 2 recombination with different alkyl radicals were measured at different pH values. We have demonstrated that the pH regulation of the controlled regime of polymerization is possible using different types of hydrophilic monomers in aqueous solution at 90 °C. It is also shown that the pH variation can be used for controlling the polymer chain length. 5559

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

ARTICLE

Scheme 2. Reaction Scheme for Alkoxyamines Preparation

’ RESULTS AND DISCUSSION Nitroxides and Alkoxyamines Synthesis. The alkoxyamines 3, 4, 5, and 6 and the nitroxide 2 were prepared according to Scheme 2. It should be noted that the alkoxyamines 3 and 9 coordinate copper ions, and therefore the solutions of 3 obtained using the standard Matyjaszewski’s procedure were dark-green due to the copper complexes. The complexes were stable in chromatography columns, so the treatment with diethylene triamine pentaacetic acid (DTPA) was needed to remove copper and to prepare the pure samples of 3 and 9. The following alkali hydrolysis yielded the alkoxyamines 4 and 5. NMR spectra of the alkoxyamines 3, 4, 5, and 6 show a double set of signals implying formation of a mixture of two isomeric forms; the transition between these forms is slow in the NMR time scale. The ratio of the isomers was solvent-dependent. The observed phenomena are accounted for by slow inversion at the N1 atom of the heterocycle.3437 Asymmetric centers at C2 and N1 produce two diastereomers (Chart 2), which, however, cannot be separated chromatographically due to the relatively high inversion rate. Similar effect was observed earlier for some alkoxyamines of imidazoline and imidazolidine series.27,29 AcidBase Equilibria in the Solutions of Nitroxides 1 and 2 and Alkoxyamines 46. The EPR spectrum of 1 is known to

Chart 2. Two Diastereomers with Asymmetric Centers at C2 and N1

5560

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry be sensitive to the protonation, showing a gradual increase of the observed hyperfine constant from 1.396 to 1.530 mT upon pH changes from 1 to 7.5. Analysis of the titration curve reveals two pKa values of 5.08 and 2.86, assigned to the protonation of amidine and pyridine groups, respectively (Scheme 3).25 pKa values of nitroxide 2 were obtained using the experimentally measured dependence of EPR spectra on pH shown in Figure 1. Analysis of the titration curve was performed in the same way as in ref 23 using two pKa values of 0.6 and 4.4, assigned to the protonation of amidine and carboxylate groups, respectively (Scheme 3). NMR spectra of the alkoxyamines 4, 5, and 6 in D2O also demonstrate pH dependence (see Figure 2, Scheme 4, Table 1SI in Supporting Information). The observed changes are in agreement with the equilibria shown in Scheme 4. The exchange between different protonated/deprotonated forms is fast in the NMR time scale and produces narrow or moderately broadened signals at the averaged position in the NMR spectrum. The signals of diastereomers 4A and 4B are clearly resolved. This allows for the monitoring of pH-associated changes for each isomer separately. At pH 12.5 the deprotonated form I of alkoxyamine 4 is observed in the NMR spectrum. The increase of the pH value does not lead to the spectral changes. When the solution was acidified to pH = 7 the chemical shift of the Me protons in amidine group (signal b in figure 1 a) changed from 2.8 ppm up to 3.1 ppm. Scheme 3. Protonation of Functional Groups in Nitroxides 1 and 2

ARTICLE

This indicates the protonation of amidine group. In this case the form II is predominant. The value of pKa obtained from the titration curve is 9.8. Decrease of the pH value from 7 to 4 leads to the protonation of pyridine ring and subsequent changes in the chemical shift of c and c0 signal. The value of pKa assigned to the protonation of pyridine group is 5.25. The analysis of signals in the aliphatic zone is complicated due to the presence of 8 different diastereoisomeric forms of alkoxyamine 4, and therefore no titration curve was built for the signals in this zone. It should be noted that the carboxylate group of alkoxyamine 4 is protonated at pH close to 3.8 for alkoxyamine 5 (see below). Thus, the form IV is predominant at this pH range. The titration curves for alkoxyamine 6 are similar. The values of pKa assigned to the protonation of amidine and pyridine groups were found to be 8.0 and 5.0, respectively (Figure 2e and f). Similar analysis was carried out for the titration curves of alkoxyamines 5. The value of pKa for protonation of imidazoline ring was found to be 5.7 (curve b on figure 2 d). The signal of methyl groups isobutyric moiety undergoes downfield shift upon protonation of adjacent carboxylate group (pKa = 3.7, curve a in Figure 2d). Signal d is affected by the protonation of the imidazoline ring and carboxylic group, and therefore the titration curve exhibits two values of pKa values (curve d in Figure 2d) Homolysis Rate Constant kd. The observed homolysis rates of the alkoxyamines 3, 4, 5, and 6 are pH-dependent (Figure3, Table 1) in a good agreement with the expected consequences of the consecutive protonation of their basic groups. The pHassociated changes in the kd values are observed for both polar (water) and nonpolar (chlorobenzene) solvents, and therefore the effect cannot be attributed to the variation of the polarity of the reaction medium. It should be noted that the addition of acid has no effect on the value of kd for TEMPO-based alkoxyamines in organic media.20 However, DFT calculation performed by Mazarin et al.16 showed that the protonation of alkoxyamine nitrogen may result in the increase of the activation barrier for CO bond cleavage. Presumably, that protonation does not take place within the pH region used in the present study. Figure 3 shows the kinetics of homolysis of 4 in aqueous solution at different pH and in chlorobenzene in the presence or absence of acid. Alkoxyamine 4 shows much higher homolysis rate constants kd at pH 8.7 when deprotonated form is predominant:

Figure 1. (a) EPR spectra of nitroxide 2 obtained at different pH values (room temperature, 104 M nitroxide aqueous solution); aN (pH 1.4) = 1.395, aN (pH 3.1) = 1.502; aN (pH 9.85) = 1.51 (b) pH-dependence of nitrogen hyperfine constant for nitroxide 2 (titration curve). 5561

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

ARTICLE

Figure 2. pH dependence of the 1H NMR spectra of alkoxyamines 4, 5, and 6 (left side) and titration curves (right side) of alkoxyamines 4, 5, and 6 obtained for 0.02 M alkoxyamine solutions in D2O at room temperature. tBuOH signal was used as reference.

at the same temperature, the kd value at acidic pH is 15-fold lower than that for basic medium (Table 1). Table 1 summarizes the results of the kd measurements for alkoxyamines 36 in media of different acidity. For comparison, the activation energy of

homolysis Ea was estimated using the averaged frequency factor A = 2.4  1014 s1 (Table 1).38 One can see that all alkoxyamines exhibit pH-dependence of the observed values of kd. Indeed, kd measurements at pH = 11 in water, when the form I of the 5562

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

ARTICLE

Scheme 4. Forms of Alkoxyamines 36 at Different pH

alkoxyamine 4 is predominant, give the highest kd values. Acidification to pH = 8.7 leads to a decrease in kd value due to the predominance of form II. In pure chlorobenzene solution alkoxyamine 4 is expected to be in the form II, which is confirmed by the close values of kd observed in water at pH = 8.7 and in pure chlorobenzene solution. The addition of 1 or 2 equiv of CF3COOH to the solution of 4 in chlorobenzene gives similar results, indicating that the acidity of the media may lead to the protonation of amidine, while the pyridine nitrogen presumably remains deprotonated (form III). The addition of 4 mM concentration of CF3COOH leads to further decrease in kd value, which is obviously caused by alkoxyamine dication (form IV) predominance (Figure 3b, Table 1). The protonation of the nitroxide fragment results in the decrease of kd by a factor four: namely such effect was observed for 3 and 6 (Table 1), for which the protonation of the nitroxide fragment is unambiguous. A 15fold effect is observed for 4, but that likely involves synergetic effect due to the protonation of both alkyl and nitroxide fragments. It should be mentioned that in the organic solution pure alkoxyamine 4 exists in the zwitterionic form, which is indicated by good agreement between the values of kd for pure 4 in C6D5Cl and 4 in water at pH = 8.7. It is known that diastereomers may exhibit the same or different kd values.39 In our experiments, the integration of several NMR lines corresponding to different diastereomers was used to determine the kd values. The difference observed

for different diastereomers was within the range of the experimental error at all pH values. According to the literature data, the values of the rate constant kd are usually similar for alkoxyamines based on carboxylic acid and carboxylate alkyl fragment40 unless the case hyperconjugation effect takes place. Since protonation of carboxylic and pyridine groups occurs at the close pH values, the hyperconjugation effect cannot be separated clearly from the electron withdrawing effect. A similar difference in kd values for the protonated and deprotonated forms was observed for alkoxyamine 5 and 6. Unlike 4, alkoxyamines 5 and 6 are poorly soluble in benzene. Therefore the kd values were measured for aqueous solutions only. In basic solution alkoxyamines 5 and 6 showed larger kd than in acidic media, demonstrating the same trend as for alkoxyamines 3 and 4. It should be mentioned that the values of kd for the protonated/deprotonated forms of 6 are at least 10 times smaller than those for alkoxyamine 4, indicating the influence of the amide residue on the homolysis of CO bound. Note, that the possible protonation of NH2 group should be excluded as it would increase the value of kd as observed by Bremond et al.21 Recombination Rate Constant kc. The rate constants kc of the radical cross-coupling were measured by LFP method. Alkyl radicals were generated by photolysis of the corresponding symmetric ketones, and their decay was monitored at the absorption maximum of 315 nm.41 In the absence of nitroxides, 5563

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

ARTICLE

Figure 3. Kinetics of homolysis of alkoxyamine 4 (a), alkoxyamine 5 (c), and alkoxyamine 6 (d) in D2O (0.02 M solution) at 368 K measured by 1H NMR in the presence of 40 equiv of ascorbic acid/ascorbate at different pH. Kinetics of homolysis of alkoxyamine 4 (b) in chlorobenzene (104 M solution) at 373 K measured by ESR in the presence of oxygen as a scavenger. Experimental conditions are presented in the figure.

Table 1. Observed Dissociation Rate Constants kd and Evaluated Activation Energies of Dissociation Ea for Alkoxyamines 36a alkox

method

predominant form

3

EPR

III

NMR

4

EPR

NMR

NMR

6

a

NMR

solvent

T, K

conditions

kd, 104 s1

Ea ( 0.5, kJ mol1

4  103 M of CF3COOH

373

7.2 ( 0.3

125.0

II

103 M of CF3COOH

373

10 ( 0.1

124.0

I

pure solution

373

27 ( 3

121.0

I

treated with NaOH

373

30 ( 3

120.5 129.0

C6H5Cl

pH = 3.0

363

0.5 ( 0.05

II

pH = 6.7

363

3.0 ( 0.5

124.0

I

pH = 8.7

363

5.0 ( 0.5

123.0

4  103 M of CF3COOH C6H5Cl, 104 M of CF3COOH

373 373

3.4 ( 0.1 17 ( 1

127.4 122.0

III

C6H5Cl, 2  104 M of CF3COOH

373

18 ( 1

122.0

II

C6H5Cl, pure

373

43 ( 5

119.5

II

C6H5Cl, treated with NaOH

373

52 ( 3

119.0

IV

pH = 3.0

368

26 ( 0.1

126.5

III

pH = 6.7

368

12 ( 0.5

121.8

III

IV III

D2O/CD3OD = 1:1

C6H5Cl

II

pH = 8.7

368

35 ( 2

118.5

I III

pH = 11 pH = 3.0

368 368

45 ( 2 1.2 ( 0.2

117.8 128.9

II

pH = 5.2

368

4.2 ( 0.2

125.0

I

pH = 9.5

368

14 ( 0.5

121.4

pH = 1.5

363

0.2 ( 0.02

132.5

pH = 3.2

363

0.2 ( 0.02

132.5

II

pH = 7.0

363

0.55 ( 0.01

129.5

I

pH = 10.0

363

1.3 ( 0.1

126.0

IV

D2O

Concentration of alkoxyamine was 20 mM for NMR experiments and 1 mM in EPR experiments.

the carbon-centered radicals decay in the reactions of dimerization and disproportionation by second-order kinetics with the parameter kt/ε = 7  106 cm s1 for methyl propionyl radical

(MP) and kt/ε = 6  106 cm s1 for t-Bu-isobutiryl radical (tBiB), where ε is the extinction coefficient of the corresponding alkyl radicals. In the presence of nitroxides, the decay becomes 5564

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

ARTICLE

Figure 4. Dependence of the pseudo-first-order decay rate constant kobs of alkyl radicals tBiB (a) and MP (b) on the concentration of nitroxides 1 and 2 at room temperature, observed during the photolysis of PEST and MEST ketones. Acetonitrile/water mixture (1:1): (1) 1 in completely protonated form III; (r) 1 in deprotonated form I; (9) 2 in completely protonated form III; (0) 2 in deprotonated form I.

Table 2. pH Dependencies of the Rate Constant kc of CrossCoupling of Nitroxides 1 and 2 with MP and tBiB Alkyl Radicals entry 1•

2•

kc, 108 M1 c1

pH

MP•

1

5.3 ( 2

2

tBiB•

2 3

9.3 ( 3 4.2 ( 0.7

9.2 2

4

1.6 ( 0.4

5

2.6 ( 0.4

0.5

6

1.4 ( 0.3

9

tBiB•

10

exponential, the observed pseudo-first-order rate constant kobs is proportional to the nitroxide concentration kobs = k0 þ kc[NO•] (Figure 4). It was shown that the values of cross-coupling rate for constants of different nitroxides kc are markedly below the diffusioncontrolled limit of about 5  109 M1 s1 and only slightly depend on temperature, showing very small, or even negative, activation energy.19,42 In general, radical coupling reactions exhibit very low (below 10 kJ mol1) Ea values. Therefore, the variations of kc with T are weak. This indicates the entropy control of the reaction, i.e., the hindered approach of the alkyl radical to the oxygen-centered radical. Thus, the measurements of kc were performed at room temperature, and kc values at T ≈ 90100 °C were assumed to be approximately the same. The values of kc were obtained for the nitroxyl radicals 1 and 2 at different pH (Table 2, Figure 3). It should be noted that the absolute values of kc are close to the ones reported for such types of nitroxides, i.e., 1.08.0  108 M1 s1.19 Figure 4 and Table 2 demonstrate that the values of kc for the protonated forms of nitroxides 1 and 2 and tBiB• are approximately two times larger than for the deprotonated forms. Thus, the observed influence of protonation on kc is not large, and it is opposite to that on kd. The exception is the case of nitroxide 1 and MP• alkyl radical, where the value of kc for the protonated state is slightly lower than for the deproponated state, but it should be mentioned that in this case the values of kc are close to diffusion limit, which causes large experimental error. It was previously shown44,43 that kc is

influenced by the same effects as kd, i.e., stabilization, steric, and polar effects, and they are expected to be anticorrelated to those observed for kd. The effect of polar group on kc was previously clearly observed for the several nitroxides, but the difference in kc values due to the polar effect is not large.44 Thus, the observed results are in a good agreement with our expectations based on the literature data and the structural changes in nitroxides upon protonation. Thermal Stability of Nitroxides. The thermal instability of a nitroxide mediator can sometimes destroy the controlled regime of polymerization.29 The presence of acid in the reaction medium can favor the decomposition of nitroxides.45 Thus, the thermal stability of nitroxides 1 and 2 at different pH (in acidic and basic media) has to be verified. The measurements were performed for 104 M degassed solutions of nitroxides 1 and 2 in chlorobenzene in the presence of 120 equiv of CF3COOH or excess of base. The EPR signals of nitroxides were monitored at 363 K for 36 h. No changes in the intensities of EPR signals were observed, so one can conclude that the imidazoline nitroxides 1 and 2 are sufficiently stable under acidic and basic conditions. Polymerization. Taking into account the values of alkoxyamine homolysis rate constants kd and of nitroxide and alkyl radical recombination rate constants kc, one can predict the controlled and “living” regimes of polymerization of particular monomer using a Fischer diagram.46 Figure 5 shows the diagram for the polymerization of styrene (Sty) at 140 °C and acrylamide (AAm) at 90 °C. Since the reactivities of monomers and polymers differ, the values of the rate constants kd and kc measured for monomers in the Fisher diagram need to be adjusted. Typically, the activation energy of homolysis for imidazoline-based alkoxyamines with the isobutyrate alkyl fragment is by 5 kJ mol1 smaller than that with the ethylbenzyl fragment.27 The value of kc measured for the recombination of 1 and tBiB alkyl radical was divided by 30 in agreement with the literature data.47 Since the chain length effect for styrene was found to be negligible,4850 the rate constants kd(Sty) and kc(Sty) were used without further adjustment. In the construction of Fischer’s diagram for acrylamide the values of kc for alkoxyamine 6 and the values for nitroxide 1 and MP alkyl radical recombination were taken. To take into account the chain length 5565

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

ARTICLE

Figure 5. Fischer diagrams46 for the polymerization of monomers (a) Sty at 140 °C controlled by nitroxides I (with the correction of rate constants), the monomer to initiator ratio is 1000/1, kp = 103 M1 s1, kt = 1.7 0  108 M1 s1;54 (b) AAm at 90 °C controlled by nitroxides 1:, the monomer to initiator ratio is 1000/1, kp = 6.6  103 M1 s1, kt = 1.6  108 M1 s1.55 The gray region corresponds to the controlled and “living” polymerization. In both diagrams polymerization time corresponds to 90% of conversion.

and the penultimate unit effect on the values of kd and kc, it was assumed that in the pAAm alkyl chain the value of kd is 2 time greater51 and the value of kc is 10 times smaller52 than it was measured for the alkyl radical, as observed with TEMPO-based macro-alkoxyamines. Since there is no literature data on the chain length and penultimate effect for AAm polymerization, the Fischer’s diagram can be used for the qualitative evaluation of the influence of protonation on polymerization. Two main challenges were addressed in these experiments: (i) polymerization temperature below 100 °C for the polymerization in aqueous media and (ii) relatively high molecular weight of the final polymer (∼105 g mol1 at high monomer conversion). These two factors shrink the controlled region in the Fischer’s diagram making the achievement of the controlled regime rather complicated. Figure 5a shows that one can expect the controlled polymerization of Sty mediated by forms IN1 and IIN1 (Scheme 3) of the nitroxide 1. The Sty polymerization mediated by the protonated form IIIN1 (Scheme 3) of the nitroxide 1 is expected to be slow. Figure 5b shows that the polymerization of AAm mediated by the nitroxide 1 in the protonated form IIIN1 should be controlled and living, while the polymerization mediated by the nitroxide 1 in the deprotonated form IN1 is too slow and, consequently, not controlled. Figure 6 and Table 2SI in Supporting Information demonstrated the result of the Sty polymerization initiated by the different forms of alkoxyamine 4. As it is expected from the Fischer diagram analysis the polymerization of Sty initiated by the form II of alkoxyamine 4 proceeds faster than initiated by form IV (Figure 6b). In all cases the polymerization of Sty is controlled: the linear increase of Mn with conversion is observed and the PDI values for the polymers obtained are low (Figure 6b, Table 2SI in Supporting Information). In the presence of acid the polymerization proceeds slower since the homolysis of macroalkoxyamine based on the protonated form of nitroxide 1 is slow. Slow monomer conversion can be also attributed to the suppression of self-initiation process of Sty in the presence of acid as reported by Buzanowski et al.53 Slow dissociation of the protonated form of macroalkoxyamine during polymerization of Sty under acidic conditions results in the slight increase of PDI to 1.6 (Figure 6, Table 2SI, entry 3).

Figure 7a demonstrates the plot of molecular weight versus conversion for the polymerization of AAm initiated by the alkoxyamine 4 at acidic pH in aqueous solution T = 90 °C at the monomer to initiator ration 1000/1 (see also Table 3, entry 1). The polymerization of AAm was stopped at high monomer conversions (5070%) because the reaction mixture became viscous. The Mn versus conversion plot is linear, which indicates that the number of propagating radicals remains constant. The final molecular weight Mn = 3.4 0  104 g mol1 (relative to PEO standards) is in good agreement with the theoretical value Mn,th = 3.9  104 g mol1. The final dispersity index is 1.16, which is far below the theoretical limit for the noncontrolled radical polymerization. This means that the polymerization of AAm initiated by alkoxyamine 4 under acidic conditions proceeds in the controlled mode. It should be mentioned that the polymerization of AAm performed under basic conditions for 4 h does not result in the monomer conversion (Table 3, entry 2). The same influence of pH on the polymerization of AAm was observed for different monomer/initiator ratios. At acidic pH, controlled regime of polymerization was observed, that led to lower PDI value and high monomer conversion (Figure 7a and c, Table 3 entries 1, 3, 4). The controlled regime of AAm polymerization at acidic pH can occur for several reasons. First, Fischer’s diagram predicts controlled regime of polymerization only in acidic medium. Second, this might be due to some unexpected effect of pH on the intra- and intermolecular H-transfer reaction occurring with alkoxyamine. The influence of pH on the intra- and intermolecular H-transfer reaction in alkoxyamines is the topic of future investigations. The polymerization of sodium 4-vinylbenzenesulfonate (SS) initiated by the alkoxyamines 4 and 5 was also performed at different pH with the same monomer to initiator ratio and temperature. The polymerization of SS in aqueous solution at 90 °C at pH = 9.2 (Table 3, entry 8, Figure 8b) in the presence of alkoxyamine 4 with monomer to initiator ratio 1000/1 was characterized by a linear growth of molecular weight and decrease of dispersity with conversion. After 1 h, the monomer conversion stopped with the final value of 40%. The polymer obtained has the dispersity index 1.75. Thus, it can be concluded 5566

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

ARTICLE

Figure 6. Polymerization of Sty at 140 °C initiated by protonated/ deprotonated forms of alkoxyamine 4 and controlled by nitroxides 1. Monomer to initiator ratio is 1000/1. (a) Kinetics plot for polymerization; lines represent linear fit of the experimental data points. (b) Evolution of molecular weight and dispersity. (9) pure alkoxyamine, form II; (b) alkoxyamine in the presence of 1 equiv of CF3COOH, form III; (2) alkoxyamine in the presence of 10 equiv of CF3COOH, form IV. Solid line shows the theoretical Mn, dashed lines represent linear fit of the experimental data points.

Figure 7. (a) Kinetics of acrylamide polymerization at T = 90 °C initiated by the alkoxyamine 4 and controlled by the nitroxide 1 in form III at pH = 3.0 with different monomer to initiator ratio (M:I), lines represent linear fit of the experimental data points. (b) Evolution of molecular weight and dispersity. Dotted line is the linear fit of the experimental data points.

that the conditions of polymerization were not optimal. In the case of the monomer to initiator ratio equal to 500/1, higher conversion (65%) and lower PDI (1.32) were observed (Table 3, entry 12). When the polymerization of SS was carried out at pH = 7.3 and the monomer to initiator ratio 1000/1 (Table 3, entry 10, Figure 8b), 70% of monomer conversion was reached in 3 h. A linear growth of molecular weight was observed up to ca. 50% of the monomer conversion. Later on, the rate of the molecular weight growth decreases, which can probably be attributed to the viscosity increase. Nevertheless, the final dispersity index of the polymer obtained was 1.15, which

is far below the theoretical limit of 1.5 for the controlled radical polymerization. This demonstrates that the acceptable conditions for the controlled regime of polymerization were achieved. It should be noted that the self-initiation process for SS at 90 °C is negligible and cannot affect the kinetics of polymerization. The results obtained can be compared with the data reported by Phan et al.32 and Nicolay et al.33 The authors of these works described NMP of SS up to 63% of conversion and low dispersity, but the molecular weights of the polymer were smaller. Under acidic conditions, the evolution of Mn with 5567

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

ARTICLE

Table 3. Experimental Conditions and Characteristics of the Polymers Obtained for the Polymerization of Hydrophilic Monomers at 90 °C Initiated by Alkoxyamines in Different Forms

entry initiator monomer 1 2 3 4 5 6 7 8 9 10 11 12 13 14 a

4 4 4 4 5 5 5 4 4 4 4 4 5 5

AAm AAm AAm AAm AAm AAm AAm SS SS SS SS SS SS SS

M:I

conversion pH % PDI

1000/1 3.2 1000/1 9.9 250/1 3.0 100/1 3.0 1000/1 5.2 1000/1 2.7 500/1 6.5 1000/1 9.2 1000/1 6.7 1000/1 7.3 1000/1 3.0 500/1 10.1 1000/1 6.3 1000/1 2.5

Mna 105 g mol1

Mn,th 105 g mol1

55.7 0 76.0 55.0 0 20.0

1.16

0.34

0.40

1.20 1.80

0.72 0.21

0.8 0.20

4.0

2.0

0.14

40.0 51.7 70.0 80.0 65 35 35

1.75 1.2 1.15 1.2 1.32 1.80 1.60

1.5 1.2 1.37 2.0 1.0 1.1 4.3

0.82 1.07 1.40 1.70 0.70 0.72 0.72

Relative to PEO standards.

conversion is linear for the polymerization of SS initiated by the alkoxyamine 4 (Table 3, entry 11, Figure 6b), though the molecular weight of the polymer is slightly higher than the calculated values. The polymerization was controlled up to 80% of the monomer conversion with the low dispersity of the polymer obtained. The livingness of the polymer was verified by reinitiation experiments. Poly-Sty obtained in the experiments 13 (Table 3SI in Supporting Information) was precipitated in cold methanol and used for the polymerization of new portion of Sty. In all cases the increase of molecular weight and the absence of macroinitiator traces were observed by means of GPC indicating the living character of the polymerization (Figure 9). Figure 9a presents the results of the reinitiation test for poly-Sty obtained with the form II of alkoxyamine 4 (Table 2SI in Supporting Information, entry 1). The traces of the low molecular weight poly-Sty observed in GPC can probably be due to the self-initiation process at 140 °C. In the presence of acid selfinitiation of polymerization of styrene is suppressed,53 and thus the low molecular weight peak is absent in Figure 9b. The analysis of experimental results taking into account three contributions (polymer formed from macroinitiator, dead polymer chains and low molecular weight polymers determined by self-initiation of styrene) allows us to evaluate the fraction of living chains to be larger than 85%. Thus, a higher molecular weight than expected (Figure 8b) could be explained by some loss of living ends. The poly-(SS) obtained by the polymerization in the presence of alkoxyamine 4 under neutral conditions at 70% of conversion (Table 3, entry 10) was precipitated in cold methanol, dried in vacuum, and used as a high molecular weight initiator for the polymerization of AAm. Acidic conditions and temperature 90 °C were chosen for the polymerization of AAm. Under these conditions, 69% of AAm conversion were achieved with the final molecular weight Mn = 4.3  105 g mol1 and dispersity 1.24. No traces of poly-(SS) were observed in the final reaction mixture. This indicates that the polymerization of SS in the presence of alkoxyamine 4 is living. The evaluated fraction of the living chain was higher than 95%. The GPC traces obtained for macroinitiator and co-polymer are presented in Figure 9.

Figure 8. (a) Kinetics of 4-vinylbenzenesulfonate polymerization at 90 °C initiated by the different forms of alkoxyamine 4 and controlled by the nitroxide I, monomer to initiator ratio 1000/1, at pH = 9.2 (9, form I of nitroxide 1), pH = 7.3 (b, form II of nitroxide 1), and pH = 3.0 (2, form III of nitroxide 1), lines represent linear fit of the experimental data points. (b) Evolution of molecular weight and dispersity. Solid line shows the theoretical Mn, dotted line is the linear fit of the experimental data points.

The polymerization of both AAm and SS initiated by alkoxyamine 5 and controlled by nitroxides 1 in acidic, basic or neutral media was not successful (Table 3, entries 57 and 13, 14, Figure 2SI in Supporting Information). Polymerization of AAm was uncontrolled giving low monomer conversion (∼20%) with high dispersity index of the polymer obtained. In the polymerization of SS initiated by alkoxyamine 5, the molecular weights deviated from the calculated values, the dispersity index was above 1.5, and only low monomer conversions could be achieved. The Mn versus conversion and PDI versus conversion plots are presented as Supporting Information. 5568

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

ARTICLE

Figure 9. Normalized GPC traces of (a) poly-Sty (entry 1, Table 2SI in Supporting Information) (dashed line) and poly-SS obtained in reinitiation experiment (solid line); (b) poly-(SS) (entry 10, Table 3) (dashed line) and poly-(SS)-b-poly-(AAm) (after reinitiation experiment) (solid line).

’ CONCLUSION Imidazoline alkoxyamines with functional groups containing multiple ionizable centers were synthesized to quantify the influence of reversible protonation on the dissociation rate constant kd and on the cross-coupling rate constants of nitroxide and alkyl radicals kc. Using the polymerization of three representative monomers (styrene, alcrylamide, and styrene sulfonate) mediated by nitroxides with protonable groups, we have clearly demonstrated that the pH value crucially affects the polymerization regime, changing it from the controlled to the uncontrolled mode. In the case of the controlled radical polymerization, the pH dependence of the polymer length chain was observed. It is shown that the pH-induced switching of kd and kc opens an easy and convenient way for the preparation of block co-polymers via NMP since it allows the achievement of controlled regime of polymerization for different monomers at the same temperatures and monomer to initiator ratios. Note that alkoxyamine 4 is wellsoluble in organic solvents and water, which makes the polymerization of different types of hydrophobic/hydrophilic monomers possible. The controlled polymerization of acrylamide and 4-styrenesulphonate initiated by 4 was observed in water. The polymerization of Sty initiated by different forms of 4 was controlled and living, affording different rates of polymerization and lengths of polymer chain depending on protonation. Successful preparation of SS-AAm block co-polymers in the controlled and living fashion in aqueous solutions requires different parameters of polymerization at different stages of the process, which cannot be achieved with the use of conventional nitroxides such as TEMPO, TIPNO, and SG1. However, the use of alkoxyamines with pH-switchable proton made it possible: the SS polymer was prepared at pH 7 (form II of 4), and the polymerization of AAm block was performed at pH 3 (form IV of 4). The polymerization proceeded in a controlled manner, and the polymers obtained were “living” making the preparation of block co-polymers possible. Thus, for the first time the concept of pH-switchable alkoxyamines was successfully applied to the controlled polymerization of hydrophobic and hydrophilic monomers in aqueous solutions at temperatures below 100 °C to obtain high molecular weight polymers and block co-polymer.

thickness was 1 mm). The 1H and 13C NMR spectra were recorded for 510% solutions using the signal of the solvent as a standard (400.134 MHz for 1H and 100.624 MHz for 13C). All chemicals were used as received. The nitroxide 1 was prepared according to the previously published procedures.25 To confirm the NMR peak assignments of alkoxyamines 37, 9 we used 1H NMR, 13C NMR J-mod spectrum and COSY 45, 200 MHz for 1 H. 1H and 13C NMR spectra of compounds 35, 9. 1H NMR spectra of alkoxyamine 3, 4, 6, 7, 9 were recorded in CDCl3; 1H NMR spectrum of alkoxyamine 5 was recorded in CDCl3/CD3OD; 13C NMR J-mod spectra of alkoxyamines 3, 4, 6, 7, 9 were recorded in CDCl3; 13C NMR J-mod spectrum of alkoxyamine 5 was recorded in CDCl3/ CD3OD; COSY 45 NMR spectra of alkoxyamines 4, 5 were recorded in D2O; COSY 45 NMR spectrum of alkoxyamine 6 was recorded in DMSO-d6.

’ EXPERIMENTAL SECTION

5,5-Diethyl-2,4-dimethyl-2-(2-methoxycarbonylethyl)2,5-dihydro-1H-imidazol-1-oxyl (8). The crystalline powder of 7

Alkoxyamines and Nitroxides Synthesis. The IR spectra were recorded in KBr pellets (the concentration was 0.25%, the pellet

(0.9 g, 3.7 mmol) was added portionwise to 0.1 M solution (100 mL) of diazomethane in ether. The suspension was stirred until a homogeneous

2-(2-Carboxyethyl)-5,5-diethyl-2,4-dimethyl-2,5-dihydro1H-imidazol-1-oxyl (2). Prepared from 3-ethyl-3-hydroxyaminopentan-2-one56 according to Scheme 2 (see Results).

3-(5,5-Diethyl-1-hydroxy-2,4-dimethyl-2,5-dihydro-1Himidazol-2-yl)propanoic Acid (7). Suspension of 3-ethyl-3hydroxyaminopentan-2-one hydrochloride (1.00 g, 5.5 mmol), levulinic acid (2.00 g, 17 mmol), and ammonium acetate (2.00 g, 26 mmol) in methanol (5 mL) was bubbled with argon and then stirred under argon for 4 days. Reaction mixture was diluted 5-fold with brine and extracted with chloroform (10 mL  3 times). The combined extract was washed with brine and dried with MgSO4. Chloroform was removed under reduced pressure. The residue was triturated with ether/hexane mixture 1:1. The precipitate was filtered off and recrystallized from ethyl acetate/hexane mixture 1:1 to yield 0.92 g (70%) of 7 as colorless crystals, mp 179183 °C (hexane/ethylacetate 1:1). Found: C, 59.42; H, 9.15; N, 11.32; calcd for C12H22N2O3: C, 59.48; H, 9.15; N, 11.56. IR (KBr), ν, cm 1: 3187 (br), 2926 (br), 1706, 1654, 1462 (br), 1375, 1303, 1263, 1220, 1169, 1144, 1107, 1038, 946 (br), 888. 1H NMR δ 400 MHz (CDCl3, CD3OD)/ppm: 0.62 and 0.67 both 3H t, J = 7 Hz (2  CH3, Et-C), 1.08 3H s (C2-CH3), 1.27 and 1.76 both 1H ABt, JAB = 15 Hz Jt 7 Hz (C2-CH2), 1.42 1H m and 1.50 3H m (2  CH2, Et-C), 1.64 3H s (C4-CH3), 2.03 and 2.11 both 1H ABt, JAB = 15 Hz Jt 7 Hz (CH2-CO). 13C NMR δ 100 MHz (CDCl3, CD3OD)/ppm: 7.46 and 9.15 (CH3, Et-C), 15.88 (CH3, C4-Me), 21.08 (CH3, C2-Me), 24.47 (CH2, CH2CO), 28.05 and 28.69 (CH2, Et-C), 35.48 (CH2, C2-CH2), 77.63 (C5), 90.00 (C2), 175.94 (COOH), 176.80 (CdN).

5569

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry solution was obtained and kept for 16 h. Then manganese dioxide (2,0 g, 23 mmol) was added and the suspension was stirred vigorously for 1 h. The precipitate of manganese oxides was filtered off, and the filtrate was evaporated under reduced pressure. The residue was separated via column chromatography on silicagel, eluent chloroform/hexane (1:1), to yield 0.8 g (85%) 8 as orange oil. Found: C, 61.42; H, 9.08; N, 10.82; calced for C13H23N2O3: C, 61.15; H, 9.08; N, 10.97. IR (neat), ν, cm1: 2967, 2879, 1739, 1638, 1458, 1437, 1380, 1290, 1197, 1175, 1073, 990, 879, 845, 785, 747.

2-(2-Carboxyethyl)-5,5-diethyl-2,4-dimethyl-2,5-dihydro1H-imidazol-1-oxyl (2). Aqueous solution of NaOH 20% (1.0 mL) was added to the solution of nitroxide 2 (200 mg, 0.78 mmol) in ethanol (0.5 mL) and the reaction mixture was kept for 6 h at þ20 °C. Ethanol was evaporated under reduced pressure, then water (3 mL) and diethyl ether (5 mL) were added. The mixture was acidified to pH 3 with NaHSO4 upon stirring, organic layer was separated and dried with Na2SO4. Ether was evaporated under reduced pressure and the residue was recrystallized from ethylacetate/hexane mixture 1: 2 to give 2 as yellow crystals, mp 147148 °C. Found: C, 59.51; H, 8.63; N, 11.61; calcd for C12H21N2O3: C, 59.73; H, 8.77; N, 11.61. IR (KBr), ν, cm1: 3426 (br), 2971, 2934, 2884, 2746, 2680, 2539 (br), 1719, 1641, 1454, 1423, 1388, 1296, 1200, 1160, 1111, 1033, 947, 888.

Ethyl 2-(4-(Dimethylamino)-2-ethyl-5,5-dimethyl-2-(pyridin4-yl)-2,5-dihydro-1H-imidazol-1-yloxy)-2-methylpropanoate (3). Prepared according to Matyjaszewski’s method.57 A mixture of 2-(4-(dimethylamino)-2-ethyl-5,5-dimethyl-2-(pyridin-4-yl)-2,5-dihydro-1H-imidazol-1-oxyl (1)25 (300 mg, 1.15 mmol), 2-bromo-2-methylpropionic acid ethyl ester (300 mg, 1.54 mmol), Cu powder (350 mg, 5.52 mmol), 4,40 -di-tert-butyl-2,20 -bipyridine (12 mg, 0.045 mmol), Cu(OTf)2 (12 mg, 0.033 mmol), and benzene (5 mL) was placed in a Schlenk flask and degassed by three freezepumpthaw cycles. The solution was stirred for 5 h at 60 °C. The benzene was removed under reduced pressure, the residue was treated with ethyl acetate, and the inorganic precipitate was filtered off. The solution was carefully washed with basified 10% solution of diethylene triamine pentaacetic acid (DTPA) potassium salt (pH 8) in water to remove copper and dried with Na2SO4, and the solvent was removed under reduced pressure. The residue was separated using column chromatography on silicagel, eluent chloroform, to yield 3 (346 mg, 80%) as colorless crystals, mp 6770 °C (hexane). Found: C, 64.21; H, 8.57; N, 15.03. Calcd for C20H32N4O3: C, 63.80; H, 8.57; N, 14.88. IR (KBr), ν, cm1: 3083, 2982, 2931, 2866, 1728, 1606, 1593, 1466, 1447, 1409, 1399, 1379, 1364, 1291, 1228, 1176, 1145, 1107, 1071, 1029, 981, 910, 812, 652. 1H NMR δ 400 MHz (CDCl3), isomer A (60%) 0.83 3H t, J = 7 Hz (CH3, Et-C), 1.08, 1.31, 1.37, and 1.53 each 3H s ((CH3)2C-N and (CH3)2C-O), 1.22 3H t, J = 7 Hz (CH3, Et-O), 1.81 and 2.08 each 1H ABq, JAB = 14 Hz Jq 7 Hz (CH2, Et-C), 3.00 6H s ((CH3)2N), 4.09 2H m (CH2, Et-O), 7.30 and 8.45 each 2H AA0 BB0 (Py); isomer B (40%) 0.90 3H t, J = 7 Hz (CH3, Et-C), 0.94, 1.22, 1.29, and 1.50 each 3H s ((CH3)2C-N and (CH3)2C-O), 1.20 3H t, J = 7 Hz (CH3, Et-O), 1.85 and 2.38 each 1H ABq, JAB = 14 Hz, Jq = 7 Hz (CH2, Et-C), 2.95 6H s ((CH3)2N), 4.02 2H q, J = 7 Hz (CH2, EtO), 7.30 and 8.45 each 2H AA0 BB0 (Py). 13C NMR δ 100 MHz (CDCl3), isomer A 7.94 (CH3, Et-C), 13.87 (CH3, Et-O), 19.11, 23.32, 26.21, and 27.03 ((CH3)2C-N and (CH3)2C-O), 31.76 (CH2, Et-C), 38.75 (N-CH3), 60.77 (OCH2), 67.50 (CMe2), 81.87 (OCMe2), 89.76 (N-C-N), 123.59 (C3-Py), 148.38 (C2-Py), 152.28 (C4Py), 168.11 (CdO), 174.21 (CdN); isomer B 9.19 (CH3, Et-C), 13.87 (CH3, Et-O), 20.60, 23.55, 25.28, and 27.81 ((CH3)2C-N and (CH3)2CO), 27.52 (CH2, Et-C), 38.93 (NCH3), 60.66 (OCH2), 68.65 (CMe2), 81.50 (O-CMe2), 91.39 (N-C-N), 122.41 (C3-Py), 149.10 (C2-Py), 155.21 (C4-Py), 167.63 (CdO), 174.23 (CdN).

2-(4-(Dimethylamino)-2-ethyl-5,5-dimethyl-2-(pyridin4-yl)-2,5-dihydroimidazol-1-yloxy)-2-methylpropanoic Acid (4). A solution of NaOH (0.25 g, 6.25 mmol) in water (2 mL)

ARTICLE

was added to the solution of alkoxyamine 3 (376 mg, 1 mmol) in ethanol, and the mixture was stirred for 2 h and left overnight. Then the solution was evaporated under reduced pressure. The residue was triturated with ethyl acetate (5 mL), and the solution was filtered and evaporated under reduced pressure. The residue was triturated with diethyl ether, and the precipitate was filtered off and dissolved in chloroform/CCl4 1:1 mixture. Then diethyl ether was added dropwise with stirring. The crystalline precipitate was filtered off and washed with dry diethyl ether to give 4 (300 mg, 63%, zwitterionic form, crystalline solvate with chloroform) as colorless crystals, mp 8689. Found: C, 48.40; H, 6.18; N, 12.00, Cl 22.57. Calcd for C18H28N4O3 3 CHCl3: C, 48.78; H, 6.25; N, 11.98, Cl, 22.74. IR (KBr), ν, cm1: 3431 br, 2982, 2939, 2876, 1601 br, 1467, 1447, 1403, 1365, 1289, 1225, 1180, 1144, 1068, 1015, 815, 789, 910, 764. 1H NMR δ 400 MHz (CDCl3), isomer A (60%) 0.82 3H t, J = 7 Hz (CH3, Et-C), 1.00 3H s, 1.32 6H s and 1.40 3H s ((CH3)2C-N and (CH3)2C-O), 1.77 and 2.11 each 1H ABq, JAB = 14 Hz, Jq = 7 Hz (CH2, Et-C), 2.99 6H s ((CH3)2N), 7.38 and 8.37 each 2H AA0 BB0 (Py); isomer B (40%) 0.84 3H t, J = 7 Hz (CH3, Et-C), 0.99, 1.03, 1.27and 1.44 each 3H s ((CH3)2C-N and (CH3)2C-O), 1.89 and 2.38 each 1H ABq, JAB = 14 Hz, Jq = 7 Hz (CH2, Et-C), 2.95 6H s ((CH3)2N), 7.33 and 8.37 each 2H AA0 BB0 (Py). 13C NMR δ 100 MHz (CDCl3), isomer A 8.00 (CH3, Et-C), 19.10, 24.49, 26.20, and 27.02 ((CH3)2C-N and (CH3)2CO), 31.74 (CH2, Et-C), 38.78 (N-CH3), 67.31 (CMe2), 82.80 (OCMe2), 89.25 (N-C-N), 124.05 (C3-Py), 147.36 (C2-Py), 153.71 (C4Py), 168.38 (CdO), 179.55 (CdN); isomer B 9.28 (CH3, Et-C), 20.86, 24.90, 25.42, and 27.77 ((CH3)2C-N and (CH3)2C-O), 27.72 (CH2, EtC), 39.05 (NCH3), 68.79 (CMe2), 82.31 (O-CMe2), 91.25 (N-C-N), 122.71 (C3-Py), 148.09 (C2-Py), 156.27 (C4-Py), 167.89 (CdO), 179.78 (CdN).

Sodium 2-(2-(2-Carboxylatoethyl)-5,5-diethyl-2,4-dimethyl-2,5-dihydro-1H-imidazol-1-yloxy)-2-methylpropanoate (5). Synthesized from 8 via alkoxyamine 9 (Scheme 2) by analogy with compounds 3 and 4.

Ethyl 2-(5,5-Diethyl-2-(2-methoxycarbonylethyl)-2,4-dimethyl-2,5-dihydro-1H-imidazol-1-yloxy)-2-methylpropanoate (9). The degassed mixture of 8, 2-bromo-2-methylpropionic acid ethyl

ester, Cu powder, 4,40 -di-tert-butyl-2,20 -bipyridine, Cu (OTf)2, and benzene was stirred for 5 h at 40 °C. After processing the reaction mixture the residue was separated using column chromatography on silicagel, eluent chloroform, to give 9 in 72% yield as a colorless oil, a mixture of diastereomers. Found: C, 62.19; H, 9.27; N, 7.56; calcd for C19H34N2O5: C, 61.60; H, 9.25; N, 7.56. IR (neat), ν, cm1: 3455 (br), 2984, 2950, 2876, 1739 (br), 1667, 1451, 1437, 1378, 1285, 1231, 1174, 1145, 1102, 1027, 994, 953, 909, 887, 877, 838, 767, 685, 620, 596, 565. 1H NMR δ 300 MHz (CDCl3)/ppm: isomer A (70%): 0.71 and 0.79 both 3H t, J = 7 Hz (2  CH3, Et-C), 1.21 3H t, J = 7 Hz (CH3, Et-O), 1.29 3H s (CH3, C2-Me), 1.32 6H s (2  CH3, OC(Me)2), 1.081.58 3H m and 1.672.03 3H m (CH2, Et-C, CH2-C2), 2.29 and 2.50 each 1H m (CH2, CH2-CdO), 3.56 3H s (CH3, Me-O), 4.08 2H q, J = 7 Hz (CH2O); isomer B (30%): 0.65 and 0.80 both 3H t, J = 7 Hz (2  CH3, Et), 1.21 3H t, J = 7 Hz (CH3, OEt), 1.31 3H s (CH3, Et-O), 1.33 6H s (2  CH3, OC(Me)2), 1.081.58 and 1.672.03 6H in all both m (CH2, Et-C, CH2-C2), 2.102.27 and 2.362.58 each 1H m (CH2, CH2-CdO), 3.58 3H s (CH3, Me-O), 4.06 2H q, J = 7 Hz (CH2O). 13C NMR δ 75 MHz (CDCl3)/ppm: isomer A: 9.14, 10.54 (CH3, Et), 13.69 (CH3, OEt), 16.62 (CH3, Me-Cd), 21.06 (CH3, Me-C2), 24.28 and 24.66 (CH3, C(Me)2CO), 27.19 and 29.12 (CH2, Et), 30.02 (CH2, CH2-CdO), 37.47 (CH2, CH2-C2), 51.12 (CH3, Me-O), 60.59 (CH2, OEt-O), 80.88 (C, C(Me)2CO), 81.14 (C5), 93.27 (C2), 171.68 (CdN), 173.73 and 174.21 (COOMe, COOEt); isomer B: 8.85, 10.64 (CH3, Et), 13.69 (CH3, OEt), 16.42 (CH3, Me-Cd), 24.28 and 24.47 (CH3, C(Me)2CO), 26.14 (CH3, Me-C2), 27.14 and 29.37 (CH2, Et), 30.29 (CH2, CH2-CdO), 31.84 (CH2, CH2-C2), 51.15 (CH3, Me-O), 60.50 (CH2, Et-O), 81.04 (C, C(Me)2CO), 81.84 (C5), 93.21 (C2), 171.32 (CdN), 173.67 and 174.35 (COOMe, COOEt). 5570

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry Sodium 2-(2-(2-Carboxylatoethyl)-5,5-diethyl-2,4-dimethyl2,5-dihydro-1H-imidazol-1-yloxy)-2-methylpropanoate (5). A solution of NaOH (0.5 g, 12.50 mmol) in water (10 mL) was added to the solution of alkoxyamine 9 (627 mg, 1.69 mmol) in ethanol (10 mL), and the reaction mixture was kept at 5 °C overnight. Then the mixture was evaporated under reduced pressure. The residue was diluted with isopropanol (15 mL) and bubbled with CO2 to pH ∼6. The solution was filtered and evaporated again. The residual glassy solid was triturated with diethyl ether, filtered off, and dried in vacuum to give 5 (567 mg, 90%) as a colorless powder. Found: C, 51.49; H, 7.10; N, 7.14; Na, 12.20; calcd for C16H26N2Na2O5: C, 51.61; H, 7.04; N, 7.52; Na, 12.35. IR (KBr), ν, cm1: 3429 (br), 2969, 2943, 2879, 1702, 1662, 1581, 1456, 1398, 1378, 1358, 1301, 1230, 1187, 1154, 1104, 1009, 984, 951, 886, 833, 787, 685, 616, 596, 574, 487, 430. 1H NMR δ 300 MHz (CDCl3, CD3OD)/ppm: isomer A (70%): 0.85 and 0.93 both 3H t, J = 7 Hz (2  CH3, Et), 1.36, 1.41, and 1.47 each 3H s (C2-CH3, (CH3)2CO), 1.521.69 3H m and 1.702.00 1H m (2  CH2, Et), 1.121.23, 2.032.16, 2.192.29 and 2.402.57 each 1H m (CH2-CH2), 1.88 3H s (CH3-Cd); isomer B (30%): 0.81 and 0.92 both 3H t, J = 7 Hz (2  CH3, Et-C), 1.37, 1.38, and 1.40 each 3H s (C2-CH3, (CH3)2CO), 1.702.00 4H m (2  CH2, Et), 1.121.23 1H, 2.032.16 1H and 2.402.57 2H each m (CH2-CH2), 1.87 3H s (CH3-Cd). 13C NMR δ 75 MHz (CDCl3, CD3OD)/ppm: isomer A: 8.51 and 9.87 (CH3, Et), 15.56 (CH3, C4-Me), 20.90 (CH3, C2-Me), 23.83 and 24.72 (CH3, OC(Me)2), 26.88 and 28.26 (CH2, Et), 31.23 (CH2, CH2CO), 37.93 (CH2, C2-CH2), 80.26 (C5), 81.23 (C-ON), 92.53 (C2), 173.74 (CdN), 178.70 and 178.93 (COO h); isomer B: 8.27 and 9.99 (CH3, Et), 15.55 (CH3, C4-Me), 23.94 and 24.58 (CH3, OC(Me)2), 25.57 (CH3, C2-Me), 26.73 and 28.70 (CH2, Et), 31.51 (CH2, CH2CO), 32.46 (CH2, C2-CH2), 80.26 (C5), 81.43 (C-ON), 92.62 (C2), 173.45 (CdN), 178.37 and 178.87 (COO h).

2-((4-(Dimethylamino)-2-ethyl-5,5-dimethyl-2-(pyridin-4yl)-2,5-dihydro-1H-imidazol-1-yl)oxy)-2-methylpropanamide (6). Prepared from the nitroxide 1 and 2-bromo-2-methylpropanamide in

analogy to the described above procedure. The reaction mixture was diluted with 25% aqueous ammonia and extracted with diethyl ether. The solvent was removed under reduced pressure, and the residue was purified using column chromatography on silicagel, eluent CHCl3/ethanol 25:2, to give 80% of 4 as colorless crystals, mp 154156 °C (CCl4). Found: C, 62.19; H, 8.18; N, 19.95. Calcd for C18H29N5O2: C, 62.22; H, 8.41; N, 20.16. IR (KBr), ν, cm1: 3312, 3181, 3044, 3019, 2974, 2928, 2876, 2795, 1686, 1593, 1472, 1423, 1407, 1380, 1356, 1288, 1232, 1146, 1113, 1069, 1018, 997, 910, 816, 652. 1H NMR δ 400 MHz (CDCl3), isomer A (60%) 0.90 3H t, J = 7 Hz (CH3, Et-C), 1.17, 1.40 each 3H s and 1.50 6H s ((CH3)2C-N and (CH3)2C-O), 1.92 and 2.20 each 1H ABq, JAB = 14 Hz, Jq = 7 Hz (CH2, Et-C), 3.04 6H s ((CH3)2N), 7.27 and 8.49 each 2H AA0 BB0 (Py); isomer B (40%) 0.95 3H t, J = 7 Hz (CH3, Et-C), 0.94, 1.23, 1.42, and 1.45 each 3H s ((CH3)2C-N and (CH3)2C-O), 1.90 and 2.32 each 1H ABq, JAB = 14 Hz, Jq = 7 Hz (CH2, Et-C), 2.99 6H s ((CH3)2N), 7.27 and 8.49 each 2H AA0 BB0 (Py). 13C NMR δ 100 MHz (CDCl3), isomer A 7.94 (CH3, Et-C), 20.43, 25.26, 25,74 and 27.61 ((CH3)2C-N and (CH3)2C-O), 31.10 (CH2, Et-C), 39.12 (N-CH3), 68.29 (CMe2), 82.69 (O-CMe2), 91.72 (N-C-N), 123.19 (C3-Py), 148.93 (C2-Py), 152.47 (C4-Py), 167.85 (CdO), 176.85 (CdN); isomer B 9.42 (CH3, Et-C), 21.27, 25.03, 25.32, and 27.93 ((CH3)2C-N and (CH3)2C-O), 27.83 (CH2, Et-C), 39.12 (NCH3), 68.92 (CMe2), 82.72 (O-CMe2), 91.85 (N-C-N), 122.30 (C3-Py), 149.20 (C2-Py), 155.01 (C4-Py), 167.17 (CdO), 177.55 (CdN). Measurement of Homolysis Rate Constant kd. Values of the homolysis rate constant kd were determined either by monitoring the concentration of nitroxide by EPR (electron paramagnetic resonance) or by monitoring the concentration of alkoxyamine by 1H NMR in the presence of nitroxyl radical scavenger/alkyl radical reducing agent during the heating of the corresponding alkoxyamines.

ARTICLE

For EPR experiments, 104 M solutions of alkoxyamine in chlorobenzene were used. To prepare solutions of the corresponding basic forms, the solutions of 4 and 5 in chlorobenzene were stirred with powder of solid NaOH for 1 h. The excess of NaOH was filtered off, and the solution was used for measurements. Oxygen dissolved in chlorobenzene was used for scavenging alkyl radicals. The kinetic measurements of alkoxyamine dissociation were performed with the use of convenient EPR X band spectrometer equipped with a standard liquid nitrogen gas flow temperature controller. The growth of the nitroxyl radical concentration obeyed the first-order kinetic law. The concentration of the nitroxide after thermolysis was measured using 104 M standard solution of nitroxide 1 or 2 in chlorobenzene. For NMR experiments, 0.02 M solution of alkoxyamine 3 or 4 in the presence of 0.4 M of ascorbic acid in D2O was used. The pH was adjusted to basic values 813 with NaOD. The kinetics of alkoxyamines homolysis were monoexponential, the parameter kd was obtained by the exponential fit of the experimental kinetic curve. The value of activation energy was evaluated with the average frequency factor A = 2.4  1014 s1.38 Measurements of Recombination Rate Constant kc. The rate constants of recombination of alkyl and nitroxyl radicals were measured by Laser Flash Photolysis (LFP) technique. The detailed description of the LFP setup was published earlier.58 Symmetric ketones di-tert-butyl 2,2,4,4-tetramethyl-3-oxopentanedioate (Pest) and dimethyl 2,4-dimethyl-3-oxopentanedioate (MEst) were synthesized according to standard procedures59 and purified by column chromatography. The concentration of ketones was 36 mM, and the concentration of nitroxides varied from 0 to 2.5 mM. Solutions of ketones with or without nitroxides in a rectangular cell with the inner dimentions 10 mm 10 mm were irradiated with a excimer 308 nm laser with pulse energy up to 100 mJ and pulse duration 1520 ns. The monitoring system included a xenon short-arc lamp connected to a high current pulser, a homemade monochromator, a photomultiplier, and a digitizer. The monitoring light, concentrated in a rectangular of 3 mm height and 1 mm width, passed through the cell along the front of the irradiated window. Thus, in all experiments the excitation optical length was 1 mm, and the monitoring optical length was 8 mm. To obtain one kinetic trace, 1520 signals were averaged. All solutions were bubbled with Ar for 15 min prior to and all the time during the experiments. Polymerization. For polymerization of styrene initiated by alkoxyamine 4 (17 mg) was dissolved in 5 mL of distilled styrene. Argon was bubbled through the mixture for 15 min, and then the reaction mixture was placed into a preheated oil bath. Polymerization was performed at 140 °C. The temperature was kept constant within 1 °C. The samples of the polymerization mixture were taken by a syringe at different time intervals from the beginning of the reaction. For polymerization of hydrophilic monomers initiated by alkoxyamine 4 or 5 (1.7 mg or 1 mg) and monomer (1 g) (sodium 4-vinylbenzenesulfonate or acrylamide) were dissolved in deionized water (4 mL) in a round-bottom flask. The pH value of the solution was adjusted by adding HCl or NaOH aqueous solutions. Argon was bubbled through the mixture for 15 min, and then the reaction mixture was placed into a preheated oil bath. Polymerization was performed at 90 °C. The temperature was kept constant within 1 °C. The samples of the polymerization mixture were taken by a syringe at different time intervals from the beginning of the reaction. Sample Analysis. The conversion of monomer was determined by 1H NMR using 50 μL samples of polymerization mixture in 550 μL of CDCl3 or D2O. The molecular weights (Mn) and polydispersities were measured by GPC using a PL-gel Mixed C column (styrene) or PL-aquagelOH Mixed C column (hydrophilic polymers) and a refractive index detector. THF or water was used as eluent at the flow rate of 1 mL min1. Calibration was performed with PSt or PEO standards. 5571

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

’ ASSOCIATED CONTENT

bS

1

H NMR data of alkoxyamine 4 in D2O at different pH, δ 200 MHz, ppm; experimental conditions and characteristics of the polymers obtained for the polymerization of styrene at 140 °C initiated by alkoxyamine 4 in different forms; evolution of molecular weight and dispersity during polymerization of SS initiated by alkoxyamine 5 at different pH; 1H and 13C NMR spectra of compounds 35, 9; 1H NMR spectra of alkoxyamines 3, 4, 6, 7, 9 recorded in CDCl3 ; 1H NMR spectrum of alkoxyamine 5 recorded in CDCl3/CD3OD; 13 C NMR J-mod spectra of alkoxyamines 3, 4, 6, 7, 9 recorded in CDCl3; 13C NMR J-mod spectrum of alkoxyamine 5 recorded in CDCl3/CD3OD; COSY 45 NMR spectra of alkoxyamine 4, 5 recorded in D2O; COSY 45 NMR spectrum of alkoxyamine 6 recorded in DMSO-d6. This material is available free of charge via the Internet at http://pubs.acs.org. Supporting Information.

’ AUTHOR INFORMATION Corresponding Author

*E-mail: [email protected]; [email protected].

’ ACKNOWLEDGMENT Russian Scientific School-7643.2010.3, Division of Chemistry and Material Science (5.1.1), Russian Federal Agency for Education (P1144) are kindly acknowledged for financial support. ’ REFERENCES (1) Solomon, D. H.; Rizzardo, E.; Cacioli, P. U.S. Patent 4,581,429; Chem. Abstr. 1985, 102, 221335q. (2) Rizzardo, E. Chem. Aust. 1987, 54, 32–33. (3) Georges, M.; Veregin, R.; Kazmaier, P.; Hamer, G. Macromolecules 1993, 26, 2987–2988. (4) Nanoscale Science and Technology; Kelsall, R. W., Hamley, I. W., Geoghegan, M., Eds.; John Wiley & Sons: Chichester, U.K., 2005. (5) Hawker, C. J.; Bosman, A. W.; Harth, E. Chem. Rev. 2001, 101, 3661–3688 and references therein. (6) Controlled/Living Radical Polymerization. Progress in ATRP, NMP and RAFT; Matyjashzewski, K., Ed.; ASC Symposium series 768; American Chemical Society: Washington, DC, 2000. (7) Advances in Controlled/Living Radical Polymerization. Progress in ATRP, NMP and RAFT; Matyjashzewski, K., Ed.; ASC Symposium series 854; American Chemical Society: Washington, DC, 2003. (8) Souaille, M.; Fischer, H. Macromolecules 2001, 34, 2830–2838. (9) Grimaldi, S.; Le Moigne, F.; Finet, J.-P.; Tordo, P.; Nicol, P.; Plechot, M. International Patent WO 96/24629, 15 August 1996. (10) Benoit, D.; Chaplinski, V.; Braslau, R.; Hawker, C. J. J. Am. Chem. Soc. 1999, 121, 3904–3920. (11) Guillaneuf, Y.; Astolfi, P.; Gigmes, D.; Marque, S.R. A.; Tordo, P.; Grecci, L.; Bertin, D. Macromolecules 2007, 40, 3107–3114. (12) Bertin, D.; Gigmes, D.; Marque, S.; Tordo, P. Chem. Soc. Rev. 2011, 40, 2189–2198. (13) Benaglia, M.; Chiefari, J.; Chong, Y.; Moad, G.; Rizzardo, E.; Thang, S. J. Am. Chem. Soc. 2009, 131, 6914–6915. (14) Benaglia, M.; Chen, M.; Chong, Y.; Moad, G.; Rizzardo, E.; Thang, S. Macromolecules 2009, 42, 9384–9386. (15) Marx, L.; Hemery, P. Polymer 2009, 50, 2752–2761. (16) Mazarin, M.; Girod, M.; Viel, S.; Phan, T. N. T.; Marque, S. R. A.; Humbel, S.; Charles, L. Macromolecules 2009, 42, 1849–1859. (17) Ficher, H. Chem. Rev. 2001, 101, 3581–3610. (18) Marque, S. R. A. J. Org. Chem. 2003, 68, 7582–7590.

ARTICLE

(19) Zubenko, D.; Tsentalovich, Yu.; Lebedeva, N.; Kirilyuk, I.; Roshchupkina, G.; Zhurko, I.; Reznikov, V.; Marque, S. R. A.; Bagryanskaya, E. J. Org. Chem. 2006, 71, 6044–6052. (20) Marque, S. R. A.; Fischer, H.; Baier, E.; Studer, A J. Org. Chem. 2001, 66, 1146–1156. (21) Bremond, P.; Marque, S. Chem. Commun. 2011, 47, 4291–4293. (22) Khramtsov, V. V. Volodarsky, L. B., Use of imidazoline nitroxides in studies of chemical reactions: ESR measurements of the concentration and reactivity of protons, thiols and nitric oxide. In Biological Magnetic Resonance; Berliner, L. J., Ed.; Plenum Press: New York, 1998; Vol. 14, p 109. (23) Kirilyuk, I. A.; Bobko, A. A.; Khramtsov, V. V.; Grigor’ev, I. A. Org. Biomol. Chem. 2005, 3, 1269–1274. (24) Potapenko, D. I.; Foster, M. A.; Lurie, D. J.; Kirilyuk, I. A.; Hutchison, J. M.; Grigor’ev, I. A.; Bagryanskaya, E. G.; Khramtsov, V. V. J. Magn. Reson. 2006, 182, 1–11. (25) Pavlovskaya, M. V.; Kirilyuk, I. A.; Shchepalov, A. A.; Grigoriev, I. A.; Grishin, D. F. Polym. Sci., Ser. B 2008, 50, 356–361. Lazarev, M. A.; Kirilyuk, I. A.; Grigor’ev, I. A.; Grishin, D. F. Poly. Sci., Ser. B 2007, 49, 224–228. (26) Zubenko., D.; Kirilyuk, I.; Roshchupkina, G.; Zhurko, I.; Reznikov, V.; Marque, S. R. A.; Bagryanskaya, E. Helv. Chim. Acta 2006, 89, 2341–2353. (27) Bagryanskaya, E.; Bertin, D.; Gigmes, D.; Kirilyuk, I.; Marque, S. R. A.; Reznikov, V.; Roshchupkina, G.; Zhurko, I.; Zubenko, D. Macromol. Chem. Phys. 2008, 209, 1345–1357. (28) Li, J.; Zhu, X.; Zhu, J.; Cheng, Zh. J. Macromol. Sci., Part A: Pure Appl. Chem. 2007, 44, 41–46. (29) Edeleva, M. V.; Kirilyuk, I. A.; Zubenko, D. P.; Zhurko, I. F.; Marque, S. R. A.; Gigmes, D.; Guillaneuf, Y.; Bagryanskaya, E. G. J. Polym. Sci., Part A: Polym. Chem. 2009, 47, 6579–6595. (30) Qiu, J.; Charleux, B.; Matyjaszewski, K. Prog. Polym. Sci. 2001, 26, 2083–2134. (31) Huang, W.; Charleux, B.; Chiarelli, R.; Marx, L.; Rassat, A.; Vairon, J.-P. Macromol. Chem. Phys. 2002, 203, 1715–1723. (32) Phan, T. N. T.; Bertin, D. Macromolecules 2008, 41, 1886–1895. (33) Nicolay, R.; Marx, L.; Hemery, P.; Matyjaszewski, K. Macromolecules 2007, 40, 6067–6075. (34) Anderson, E. J.; Corrie, J. E. T. J. Chem. Soc., Perkin Trans 2 1992, 1027–1031. (35) Kost, D.; Raban, M. J. Org. Chem. 1976, 41, 1748–1751. (36) Anderson, J. E.; Casarini, D.; Corrie, J. E. T.; Lunazzi, L. J. Chem. Soc., Perkin Trans. 2 1993, 1299–1304. (37) Busfield, W. K.; Jenkins, I. D.; Thang, S. H.; Moad, G.; Rizzardo, E.; Solomon, D. H. J. Chem. Soc., Chem. Commun. 1985, 1249–1250. (38) Fischer, H.; Kramer, A.; Marque, S. R. A.; Nesvadba, P. Macromolecules 2005, 38, 9974–9984. (39) Ananchenko, G. S.; Souaille, M.; Fischer, H.; Le Mercier, C.; Tordo., P. J Polym. Sci., Part A: Polym. Chem. 2002, 40, 3264–3283. (40) Bertin, D.; Gigmes, D.; Marque, S. R. A.; Siri, D.; Tordo, P.; Trappo, G. Chem. Phys. Chem. 2008, 9, 272–281. (41) Skene, W. G.; Scaiano, J. C. Macromolecules 2000, 33, 5065– 5072. (42) Sobek, J.; Martschke, R.; Fischer., H. J. Am. Chem. Soc. 2001, 123, 2849–2857. (43) Marque, S.; Sobek, J.; Fischer., H.; Kramer, H.; Nesvadba, P.; Wunderlich, W. Macromolecules 2003, 36, 3440–3332. (44) Fischer, H.; Marque, S. R. A.; Nesvadba, P. Helv. Chim. Acta 2006, 10, 2330–2340. (45) Veregin, R. P. N.; Odell, P. G.; Michalak, L. M.; Georges, M. K. Macromolecules 1996, 29, 4161–4163. (46) Souaille, M.; Fischer, H. Macromolecules 2000, 33, 7378–7394. (47) Zubenko, D.; Tsentalovich, Yu.; Lebedeva, N.; Kirilyuk, I.; Roshchupkina, G.; Zhurko, I.; Reznikov, V.; Marque, S. R. A.; Bagryanskaya, E. J. Org. Chem. 2006, 71, 6044–6052. (48) Goto, A.; Fukuda, T. Prog. Polym. Sci. 2004, 29, 329–385. (49) Fukuda, T; Goto, A.; Ohno, K. Macromol. Rapid Commun. 2000, 21, 151–165. 5572

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573

The Journal of Organic Chemistry

ARTICLE

(50) Chauvin, F.; Dufils, P.-E.; Gigmes, G.; Guillaneuf, Y.; Marque, S.; Tordo, P.; Bertin, D. Macromolecules 2006, 39, 5238–5250. (51) Guerret, O.; Couturier, J. L.; Chauvin, F.; El-Bouazzy H., Bertin, D.; Gigmes, D., Marque, S., Fischer, H. ; Tordo, P. Advances in Controlled/Living Radical Polymerization; ACS Symposium series 854; American Chemical Society: Washington, DC, 2003; pp 412423. (52) Guillaneuf, Y.; Bertin, D.; Castignolles, P.; Charleux, B. Macromolecules 2005, 38, 4638–4646. (53) Buzanowski, W. C.; Graham, D. J.; Priddy, D. B.; Shero, E. Polymer 1992, 33, 3055–3059. (54) Beuermann, S.; Buback, M. Prog. Polym. Sci. 2002, 27, 191–254. (55) Brandrup, J.; Immergut, E. H.; Grulke, E. A. Polymer Handbook; John Wiley & Sons: New York, 2003. Osmanov, T. 0.; Gromov, V. F.; Khomikovskii, P. M.; Abkin, A. D. Dokl. Akad. Nauk SSSR 1978, 240, 910. (56) Kirilyuk, I. A.; Bobko, A. A.; Grigor’ev, I. A.; Khramtsov, V. V. Org. Biomol. Chem. 2004, 2, 1025–1030. (57) Matyjaszewski, K.; Woodworth, B. E.; Zhang, X.; Gaynor, S. G.; Metzner, Z. Macromolecules 1998, 31, 5955–5957. (58) Molokov, I. F.; Tsentalovich, Y. P.; Yurkovskaya, A. V.; Sagdeev, R.Z. J. Photochem. Photobiol. A: Chem. 1997, 110, 159–165. (59) Knuhl, B.; Marque, S.; Fischer, H. Helv. Chim. Acta 2001, 84, 2290–2300.

5573

dx.doi.org/10.1021/jo200341m |J. Org. Chem. 2011, 76, 5558–5573